Download - Bulk atmospheric deposition of persistent organic pollutants and … · 2019. 7. 30. · multicompartmental modelling (Lammel and Stemmler 2012; Scheringer et al. 2004; Stemmler and

Transcript
  • RESEARCH ARTICLE

    Bulk atmospheric deposition of persistent organic pollutantsand polycyclic aromatic hydrocarbons in Central Europe

    Barbora Nežiková1 & Céline Degrendele1 & Pavel Čupr1 & Philipp Hohenblum2 & Wolfgang Moche2 & Roman Prokeš1 &Lenka Vaňková1 & Petr Kukučka1 & Jakub Martiník1 & Ondřej Audy1 & Petra Přibylová1 & Ivan Holoubek1 & Peter Weiss2 &Jana Klánová1 & Gerhard Lammel1,3

    Received: 5 February 2019 /Accepted: 14 May 2019 /Published online: 14 June 2019# The Author(s) 2019

    AbstractPolycyclic aromatic hydrocarbons (PAHs), polychlorinated biphenyls (PCBs) and organochlorine pesticides (OCPs) are ubiq-uitous and toxic contaminants. Their atmospheric deposition fluxes on the regional scale were quantified based on simultaneoussampling during 1 to 5 years at 1 to 6 background/rural sites in the Czech Republic and Austria. The samples were extracted andanalysed by means of gas chromatography coupled to mass spectrometry. For all seasons and sites, total deposition fluxes forΣ15PAHs ranged 23–1100 ng m

    −2 d−1, while those for Σ6PCBs and Σ12OCPs ranged 64–4400 and 410–7800 pg m−2 d−1,respectively. Fluoranthene and pyrene were the main contributors to the PAH deposition fluxes, accounting on average for 19%each, while deposition fluxes of PCBs and OCPs were dominated by PCB153 (26%) and γ-hexachlorobenzene (30%), respec-tively. The highest deposition flux of Σ15PAHs was generally found in spring, while no seasonality was found for PCBdeposition. For deposition fluxes forΣ12OCPs, no clear spatial trend was found, confirming the perception of long-lived regionalpollutants. Although most OCPs and PCBs hardly partition to the particulate phase in ambient air, on average, 42% of theirdeposition fluxes were found on filters, confirming the perception that particle deposition is more efficient than dry gaseousdeposition. Due to methodological constraints, fluxes derived from bulk deposition samplers should be understood as lowerestimates, in particular with regard to those substances which in ambient aerosols mostly partition to the particulate phase.

    Keywords Bulk atmospheric deposition . POPs . PCBs . OCPs . PAHs . Central Europe . Deposition fluxes

    Introduction

    Polycyclic aromatic hydrocarbons (PAHs) and many haloge-nated substances, such as polychlorinated biphenyls (PCBs)and organochlorine pesticides (OCPs), are ubiquitous

    contaminants in the global environment (UNEP 2003).PCBs and OCPs are bioaccumulative and persistent in envi-ronmental compartments. Moreover, PCBs, many OCPs andmany PAHs are known for their toxic properties, as somecompounds in these groups are possibly carcinogenic, muta-genic and teratogenic (Ali et al. 2014; Bansal and Kim 2015;Ludewig and Robertson 2013; Ross 2004). For example,benzo[a]pyrene (BAP) has been classified as carcinogen forhumans (group 1), while other PAHs were classified as prob-able or possible carcinogens (group 2A or 2B) by theInternational Agency for Research on Cancer (IARC 2010)and are ecotoxic. Most of these substances are now regulatedunder the auspices of international conventions for the protec-tion of the environment and human health (UNEP 2008;UNECE 1999).

    PAHs are products of incomplete combustion and haveboth anthropogenic (e.g. traffic, domestic heating) and naturalsources (e.g. crude oil and wild fires) (Dat and Chang 2017).PCBs were widely used as dielectric fluids, plasticisers and

    Responsible editor: Philippe Garrigues

    Electronic supplementary material The online version of this article(https://doi.org/10.1007/s11356-019-05464-9) contains supplementarymaterial, which is available to authorized users.

    * Gerhard [email protected]

    1 Research Centre for Toxic Compounds in the Environment, MasarykUniversity, Brno, Czech Republic

    2 Umweltbundesamt, Wien, Austria3 Multiphase Chemistry Department, Max Planck Institute for

    Chemistry, Mainz, Germany

    Environmental Science and Pollution Research (2019) 26:23429–23441https://doi.org/10.1007/s11356-019-05464-9

    http://crossmark.crossref.org/dialog/?doi=10.1007/s11356-019-05464-9&domain=pdfhttp://orcid.org/0000-0003-2313-0628https://doi.org/10.1007/s11356-019-05464-9mailto:[email protected]

  • adhesives from 1930s to 1970s (Backe et al. 2002). InCzechoslovakia, the production ended in the year 1984(Christan and Janse 2005). OCPs were used in agriculturefrom the middle of the last century and dichlorodiphenyltri-chloroethane (DDT) as vector control combatting malaria intropical countries (el Shahawi et al. 2010).

    Persistent organic pollutants (POPs) and PAHs can befound in every compartment of the environment and theirphysical and chemical characteristics allow them to betransported from one compartment to another (Cetin et al.2017; Karacik et al. 2013). As semi-volatile organic com-pounds (SOCs), they are partitioning in the atmosphere be-tween the gaseous and particulate phases. This partitioning isone of the most important factors influencing the fate in theatmosphere of these SOCs (Bidleman 1988; Keyte et al.2013), and therefore their long-range transport potential.

    Exposure of ecosystems to pollutants is dominated by at-mospheric depositions, dry and wet. Dry deposition is drivenby gravity force and diffusion (Bidleman 1988), while wetdeposition is controlled by precipitation rate and intensity(Atlas and Giam 1988; Castro-Jiménez et al. 2015; Staelenset al. 2005). It has been shown that snow is more efficientlyscavenging than rain (Lei andWania 2004;Wania et al. 1998).The effect of forests on deposition into ecosystems was stud-ied, indicating the strong influences of the surface roughness,climate parameters (e.g. wind velocity) and physicochemicalproperties of the substance (e.g. the octanol/air partition coef-ficient, Koa) (McLachlan and Horstmann 1998; Nizzetto et al.2006; Foan et al. 2012).

    Efforts have been made over the last decades to quantifydeposition fluxes for various SOCs. Fluxes for large areas oren t i r e r eg i on s h ave be en e s t ima t e d b a s ed onmulticompartmental modelling (Lammel and Stemmler2012; Scheringer et al. 2004; Stemmler and Lammel 2012)and based on freshwater sediment pollution (Meijer et al.2006). It is important to verify the modelling results by directmeasurements of atmospheric deposition. There are somestudies carried out in quantifying bulk atmospheric depositionof PAHs, PCBs and OCPs experimentally in Europe (e.g.Arellano et al. 2015; Brorström-Lundén et al. 1994; Carreraet al. 2002; Jakobi et al. 2015), while additional data are avail-able for other regions, for example the USA (Schifman andBoving 2015), China (Feng et al. 2017) and the global oceans(González-Gaya et al. 2016). Moreover, additional data forwet deposition and washout ratios are available for Europe(Shahpoury et al. 2015; Škrdlíková et al. 2011).

    In previous studies, the seasonal variations of bulk deposi-tion fluxes of PAHs have been characterised, generally show-ing higher deposition fluxes in winter (Binici et al. 2014;Birgül et al. 2011; Blanchard et al. 2007; Gocht et al. 2007).Seasonal variations of PCBs have also been studied but noclear trend has been observed (Agrell et al. 2002; Blanchardet al. 2007; Brorström-Lundén et al. 1994; Carrera et al. 2002;

    Newton et al. 2014; Teil et al. 2004). Concerning OCPs, onlyfew studies exist, mainly on the isomers of hexachlorocyclo-hexane (HCH) (Brorström-Lundén et al. 1994; Carrera et al.2002; Jakobi et al. 2015; Teil et al. 2004), while the seasonal-ity of OCP atmospheric deposition has not been addressed yetin these studies.

    The aim of this study is to provide novel data on atmo-spheric deposition of PAHs, PCBs and OCPs in CentralEurope. In particular, the seasonal and spatial variations ofthese deposition fluxes were investigated at different rural/background sites along the Czech-Austrian border in 2011–2015.

    Material and methods

    Sampling

    Total (wet and dry i.e. bulk) deposition samples were simul-taneously collected near the Czech–Austrian border at threesites in the Czech Republic, Kuchařovice (KUC), Košetice(KOS) and Churáňov (CHU), and at three sites in Austria,Wolkersdorf (WOL), Unterbergern (UNT) and Grünbach(GRU). All sites are considered as background with limitedanthropogenic sources with the exception of KUC. KUC is arural site, located in the vicinity of agricultural fields and af-fected by emissions from residential area (e.g. domesticheating) located near the sampling site (around 100 m). Amap of the sampling sites is provided in Figure S1 in theSupplementary Material.

    From September 2011 to August 2012, deposition sam-plers were deployed at each site. Additional deposition sam-ples were also collected at the same three Czech sites during2012–2015. Exact sampling periods are provided in theTable S1.

    The deposition sampler used (Čupr and Pěnkava 2012)consists of a collection vessel (250 mm diameter) made ofborosilicate glass, with a stainless steel particulate filter holderlocated at the bottom of the collection vessel. A glass columncontainingXAD-2 sorbent (Supelco, USA) is connected to thebase of the filter holder and stored in a housing with a mod-erate heater (Figure S2). The sampler is based on the samplerdeveloped for and successfully applied in the MONARPOPproject (Offenthaler et al. 2009; Jakobi et al. 2015). Both sam-plers are a modification of an earlier design (DIN 2002), im-proving sampling efficiency and making sure that exclusivelyinert materials are in contact with the sample. This new designis patented (Čupr and Pěnkava 2012). The XAD resin waspre-cleaned in Soxhlet extractor for 8 h in acetone and 8 h indichloromethane (DCM), dried overnight and stored at roomtemperature. The moderate heater was used only when theambient temperature was lower than 4 °C to prevent the for-mation of ice and to ensure that snowfall is melted

    23430 Environ Sci Pollut Res (2019) 26:23429–23441

  • immediately, such that SOC deposition with snowfall is col-lected and not lost. Therefore, this sampler allows for thesimultaneous collection of dry deposited particulate matterand gaseous compounds as well as the wet deposition.Atmospheric particles were collected on a glass microfibrefilter (GFF, 70 mm, Whatman, USA) and the dissolved phasewas collected on XAD. The sampling duration was about 3months at each site. From 2013 onwards, GFFs were changedevery month. One to three months is the common range oftotal deposition sampling of POPs (Bergknut et al. 2011;Gocht et al. 2007; Jakobi et al. 2015; Newton et al. 2014).After sampling, all samples were wrapped in aluminium foiland plastic zip lock bags and stored at − 18° C until analysis.

    Given the low deposition rates and analytical challengesof SOCs, a sampling period of 3 months is appropriate (e.g.Jakobi et al. 2015). Possible sampling artefacts arephotodegradation (photolysis, ozonation), blowoff of depos-ited particles from the surface (McLachlan 1998) andvolatilisation from sampler surfaces (funnel, GFF). The frac-tion of pollutants dissolved by rainwater (or snow melt) isnot subject to volatilisation, but irreversibly trapped in XAD.It had been evaluated using a similar sampler design (funnel,resin cartridge downstream) that up to 10% of the organicscontained in the rainwater is not retained by the sampler, butsubject to breakthrough (McLachlan and Horstmann 1998)and that up to 10% of polychlorinated dibenzo-p-dioxinsand -furans present on the surface of the sampler can be lostif the surface is not rinsed after sampling (Horstmann andMcLachlan 1997). Similarly, a previous study (funnel, filterand resin cartridge downstream; Franz et al. 1991) hasshown that 25–28% of PAHs and 26–62% of PCBsremained on the different parts of the sampler after sampling(i.e. surface, deposition walls) and could be obtained byrinsing the surface. No rinsing was done in this study.Furthermore, because of the significance of surface rough-ness (McLachlan and Horstmann 1998; Pryor et al. 2007;Glüge et al. 2015), dry particle deposition to artificial sur-faces is less efficient than to natural surfaces. Consequently,such type of samplers may underestimate the total flux ofthe target compounds to terrestrial ecosystems. Nevertheless,various advantages of such type of deposition samplingshould be emphasised: comparable sampling across sitesand seasons and, particularly, including SOC deposition re-lated to snowfall, and rather low maintenance of device.

    Sample preparation and analysis

    GFFs and XAD samples were extracted with DCM using anautomatic extractor (Büchi Extraction System, B-811,Switzerland). Surrogate recovery standards (i.e. D8-naphtha-lene, D10-phenanthrene, D12-perylene, Supelco, Merck,Germany; PCB30 and PCB185, Ultra Scientific, USA) werespiked onto each GFF and XAD before extraction. The

    extracts were then concentrated using a gentle nitrogenstream. In 2011–2012, one sampler was dedicated to PAHanalysis while the second one was for PCBs and OCPs. Forthe remaining years, PAHs, PCBs and OCPs were analysedfrom a single deposition sampler and the extracts were dividedwith 10% used for PAHs and the remaining for PCBs andOCPs. PAHs extracts were transferred to a silica columnconsisting of 1 g of anhydrous sodium sulphate and 5 g ofactivated silica and were eluted with 10 mL of n-hexane and20 mL of DCM. Both fractions were collected in the samevial. PCBs and OCPs extracts were transferred to a glass col-umn consisting of 1 g of anhydrous sodium sulphate, 0.5 g ofactivated silica, 8 g of H2SO4-modified activated silica and 1 gof activated silica and were eluted with 30 mL of DCM:n-hexane (1:1).

    The PAHs were analysed using gas chromatography 6890GC (Agilent Technologies, USA) equipped with a 60 m ×0.25 mm × 0.25 μm DB5-MS column (Agi len tTechnologies, USA) coupled to a mass spectrometer (MS5975, Agilent, USA). The temperature programme was 80°C (1 min), 15 °C min−1 to 180 °C, 5 °C min−1 to 310 °C(20 min). The inlet temperature was 280 °C. The carrier gaswas He with a flow rate of 1.5 mL min−1. The temperature ofthe transfer line was 310 °C and 320 °C for the ion source. Theused regime was selected ion monitoring. PCBs and OCPswere analysed using a 7890 GC (Agilent Technologies,USA) coupled to Waters Quattro Micro GC (Waters, USA)equipped with SGE Analytical Science HT-8 (8% Ph) column(60 m × 0.5 mm × 0.25 μm, SGE Analytical Science,Australia) coupled with MS/MS (Agilent Technologies,USA). The temperature programme was 80 °C, 40 °C min−1

    to 200 °C, 5 °C min−1 to 305 °C. The inlet temperature was280 °C. The carrier gas was helium with a flow rate of 1.5 mLmin−1. The temperature of the transfer line was 310 °C and ofthe ion source was 250 °C. The used regime was multiplereaction monitoring. The target compounds in this study were15 PAHs, acenaphthylene (ACY), acenaphthene (ACE),fluorene (FLN), phenanthrene (PHE), anthracene (ANT),fluoranthene (FLT), pyrene (PYR), benzo(a)anthracene(BAA), chrysene (CHR), benzo(b)fluoranthene (BBF),benzo(k)fluoranthene (BKF), BAP, indeno(1,2,3-cd)pyrene(IPY), dibenz(a,h)anthracene (DHA) and benzo(ghi)perylene(BPE); 6 PCBs, PCB28, PCB52, PCB101, PCB153, PCB138and PCB180; and 12 OCPs, namely 4 HCH isomers (α, β, γ,δ ) , 6 DDX compounds i .e . o ,p' - and p,p ’ -DDT,d i c h l o r o d i p h e n y l d i c h l o r o e t h e n e (DDE ) a n ddichlorodiphenyldichloroethane (DDD) and penta- and hexa-chlorobenzene (PeCB and HCB).

    QA-QC

    No field blanks were collected within this study, but two fieldblanks, each consisting of XAD and GFF, from a following,

    Environ Sci Pollut Res (2019) 26:23429–23441 23431

  • methodologically identical study, were used instead. Theseblank levels of individual PAHs, PCBs and OCPs were belowthe limit of detection or low otherwise, suggesting minor con-tamination during sampling, transport and analysis. Meanblank values with standard deviations are provided inTable S2. There were also five solvent blanks analysed forPAHs and three solvent blanks for PCBs and OCPs, whichshowed levels below the detection limit except for PHE, FLT,PYR, α-HCH and γ-HCH, which had levels lower than theinstrumental limit of quantification (iLOQ). iLOQs were de-fined from the instruments as a signal to noise ratio of ten forthe lowest point of the calibration curve and are presented inTable S3.

    The recoveries of individual samples were ranging from55.6 to 117.2% for PAHs and from 63.1 to 109.1% forPCBs and OCPs. The reported fluxes have not been adjustedfor recoveries but were blank corrected, by subtracting themean concentrations of SOCs in the field blanks. To thisend, blank values < iLOQ were replaced by 0. For derivationof temporal averages, values were replaced by iLOQ/2 when-ever the determined concentrations in samples were lowerthan iLOQ.

    Results and discussion

    Atmospheric bulk deposition flux of PAHs, PCBsand OCPs and their composition profiles

    The total deposition fluxes for 15 PAHs (Σ15PAHs) total de-position fluxes ranged at 23 to 1100 ng m−2 d−1 with an aver-age value of 190 pg m−2 d−1 for all seasons and sites investi-gated (Table S4). The here found PAH deposition fluxes arecomparable with previous studies from remote or rural sites(i.e. 38–2000 ng m−2 d−1, see Table 1), but generally lowerthan those reported from urban sites (i.e. 36–20000 ng m−2

    d−1, see Table 1). This is due to the stronger influence of PAHprimary sources (e.g. road traffic, fossil fuel burners forheating) in urban areas compared to remote or rural areas.PAHs have been effectively mitigated across Europe in recentdecades (EEA 2017). However, direct comparisons with otherstudies should be done with caution, given the different PAHsconsidered (in this study, N = 15 and in others N = 7–23).

    In our study, FLT and PYR were generally the main con-tributors to PAHs’ deposition fluxes, accounting on averagefor 19% both (Fig. 1). There is one exception at KUCwhere insome samples (from autumn 2012 to summer 2013) BBF isalso a significant contributor accounting on average for 17%of the total flux of PAHs. The possible reason is because of thedifferent nature of this site, i.e. rural while all others are back-ground, which may reflect differences in emissions. Manyother studies also reported FLT and PYR as the main contrib-utors of ΣPAHs, with non-negligible contributions from PHE

    and CHR (Birgül et al. 2011; Blanchard et al. 2007; Esen et al.2008; Fernández et al. 2003; Gambaro et al. 2009; Gocht et al.2007; Halsall et al. 1997; Ollivon et al. 2002; Rossini et al.2007).

    For all seasons and sites investigated, total depositionfluxes for 6 PCBs (Σ6PCBs) and 12 OCPs (Σ12OCPs) mea-sured were 64–4400 (average value of 400 pg m−2 d−1) and410–7800 pgm−2 d−1 (average 1900 pgm−2 d−1), respectively(Table S4). These results are within the ranges spanned byother studies in Europe (Table 1). Bulk deposition of HCHwas reported up to one order of magnitude higher fromSwitzerland in the 1990s (Table 1), in accordance with emis-sion reductions achieved (EEA 2017). The results suggest thatatmospheric deposition in the 2010s is an important pathwayof pollution transfer to ecosystems in the Central Europeanbackground.

    Wet deposition was expected to be most relevant for sub-stances partitioning to the particulate phase or gaseous, butwith significant water solubility. At KOS, a significant corre-lation (p < 0.05) between the total deposition mass flux andprecipitation amount is found only for the two HCH isomers.This test was applied only for the KOS data subset, because ofits size (N = 17, whileN = 4 orN = 8 for the other sites;N is thenumber of samples). The result supports the perception ofinfluence of water solubility/air-water equilibrium (see alsobelow, Henry coefficients of α- and γ-HCH are 0.7 and0.3 Pa m3 mol−1, respectively).

    The deposition mass fluxes of PCBs were dominated byPCB153 (26%, Fig. 1). Next main contributors were PCB28,PCB138 and PCB180, depending on locations. This is inagreement with Agrell et al. (2002), but in contradiction withother studies in Europe which reported that PCBs depositionfluxes were dominated by lower molecular weight PCBs, spe-cifically PCB28, PCB52 and PCB101 (Bergknut et al. 2011;Carrera et al. 2002; Newton et al. 2014; Teil et al. 2004). Suchspatial variation of the substance pattern upon deposition isexpected as resulting from the spatial variability of precipita-t ion and tempera ture , the la t te r in f luencing al lintercompartmental processes (Stemmler and Lammel 2012;Wania et al. 2003; Wania and Westgate 2008).

    The deposition fluxes of OCPs were generally dominatedby γ-HCH, accounting on average for 30% of Σ12OCPs, atCHU for even 52%. Second tomost contributedα-HCH to thedeposition flux of OCPs, accounting on average for 15%.From winter 2011/12 until winter 2013/14, HCH concentra-tion in air of KOS were 46% of the concentration of the DDTcompounds (Shahpoury et al. 2015), but the ratio of totaldeposition fluxes of these pollutants during this period wasFHCH/FDDX = 2.4. One decade earlier, HCH abundance inair at KOS was measured ≈ 50% higher than the one ofDDX, but more than one order of magnitude higher in rain-water collected in KOS (Holoubek et al. 2007). This compar-ison clearly points to a muchmore effective wet scavenging of

    23432 Environ Sci Pollut Res (2019) 26:23429–23441

  • Table1

    Overviewof

    bulkdepositio

    nfluxes

    ofPA

    Hs,OCPs

    andPCBsreported

    from

    EuropeandtheMediterranean.

    a Differentnumberof

    congenersweremeasured.

    bOnlyp,p’

    isom

    ersweremeasured

    Site,yearof

    sampling

    Σ7PCB

    (pgm

    −2d−

    1)

    HCB(pgm

    −2d−

    1)

    Σ4HCH

    (ngm

    −2d−

    1)

    Σ6DDX

    (pgm

    −2d−

    1)

    ΣnPAH(ngm

    −2d−

    1)

    Type

    ofsampler

    Reference

    a.Rem

    ote

    Sweden;1

    996–2005

    190–2300

    0.7–11

    (α+γ)

    40–670

    b20–390

    (n=12)

    Teflon-coatedsurface

    Hansson

    etal.2006

    Andorra,N

    orway,S

    witzerland;1

    997–1998

    38–63

    (n=23)

    Stainlesssteelb

    ucket

    Fernándezetal.2003

    Andorra,N

    orway,S

    witzerland;1

    997–1998

    1033–33,00

    47–500

    5–16

    Stainlesssteelb

    ucket

    Carrera

    etal.2002

    Italy;

    2003

    553–18,217

    Glass

    bottle

    Nizzetto

    etal.2006

    Austria,S

    pain,S

    cotland,S

    lovakia;2004–2006

    3733–16,267

    33–400

    0.6–2

    (α+γ)

    Teflon

    orstainlesssteelreservoir

    Arellano

    etal.2015

    Austria,G

    ermany,Sw

    itzerland;2

    005–10

    110–211

    2–6

    (α+γ)

    395–1586

    Isolated

    funnel-adsorbersystem

    Jakobi

    etal.2015

    Sweden;2

    006–2009

    570–10,000

    Isolated

    funnel-adsorbersystem

    Bergknutetal.2011

    Sweden;2

    009–2010

    2700–6333

    0.5–4

    (α+γ)

    Isolated

    funnel-adsorbersystem

    New

    tonetal.2014

    Austria,C

    zech

    Republic,2011–2015

    73–2030a

    29–270

    0.1–4

    48–750

    23–660

    (n=15)

    Isolated

    funnel-adsorbersystem

    Thisstudy

    b.Rural

    Sweden;1

    989–1990

    2300–4500

    70–2700

    1–40

    (α+β+γ)

    210–2000

    (n=11)

    Glass

    collector

    Brorström

    -Lundénetal.1994

    Baltic

    sea;1990–1993

    270–171,000

    PUFplugs

    Agrelletal.2002

    Germany;

    2001–2002

    334–1063

    (n=17)

    Isolated

    funnel-adsorbersystem

    Gocht

    etal.2007

    Czech

    Republic,2011–2013

    64–4400a

    47–230

    0.3–2

    490–5030

    49–1100

    (n=15)

    Isolated

    funnel-adsorbersystem

    Thisstudy

    c.Urban

    UnitedKingdom

    ;1992–1992

    1310–18,200

    (n=13)

    Glass

    vessel

    Halsalletal.1997

    France;1

    999–2000

    177–2100

    (n=14)

    Aluminium

    bottle

    Ollivonetal.2002

    France;1

    999–2003

    7671–12,0547

    36–622

    (n=14)

    Aluminium

    bottle

    Blanchard

    etal.2007

    Italy;

    2002–2004

    221–10,575

    (n=17)

    Pyrexbottle

    Rossini

    etal.2007

    Turkey;

    2005

    260–19,500

    (n=15)

    Stainlesssteelp

    otEsenetal.2008

    Italy;

    2005

    101–693

    (n=15)

    Stainlesssteelv

    essel

    Gam

    baro

    etal.2009

    Turkey;

    2006–2007

    0.2–15

    (β+γ)

    11,600

    (n=10)

    Dry

    andwetcollector

    Binicietal.2014

    Turkey;

    2008–2009

    63Stainlesssteelp

    otCindorukandTasdem

    ir2014

    Italy;

    2007–2008

    92–2432

    (n=7)

    Glass

    bottle

    Amodio

    etal.2014

    Turkey;

    2008–2009

    11,400

    (n=12)

    Dry

    andwetcollector

    Birgületal.2011

    Italy;

    2015–2016

    75–1220a

    200–650

    20–280

    (n=20)

    Isolated

    funnel-adsorbersystem

    Quetal.2019

    Environ Sci Pollut Res (2019) 26:23429–23441 23433

  • HCH vs. DDT compounds, explained by the lower Henrycoefficient − 0.7 and 0.3 Pa m3 mol−1 for α- and γ-HCH,respectively, vs. 33 and 1.1 Pa m3 mol−1 for p,p’-DDE and -DDT, respectively (298 K; Jantunen and Bidleman 2006;Shen and Wania 2005; Xiao et al. 2004). Together with thecorrelation with precipitation amounts in KOS (above), it alsosupports the perception that the total deposition flux of OCPs

    is dominated by wet deposition. A prevalence of γ-HCH andα-HCH among deposited OCPs had already been reportedand attributed to abundance in air, but also wet scavenging(Carrera et al. 2002; Newton et al. 2014; Cindoruk andTasdemir 2014; Jakobi et al. 2015). These results were con-sistent among the different sites, except for KUC where theflux of OCPs was dominated by p,p'-DDE accounting on

    (a)

    0%

    20%

    40%

    60%

    80%

    100%

    KOS KUC CHU WOL UNT GRU

    ACY ACE FLN PHE ANT FLT PYR BAA

    CHR BBF BKF BAP IPY DHA BPE(b)

    0%

    20%

    40%

    60%

    80%

    100%

    KOS KUC CHU WOL UNT GRU

    PCB28 PCB52 PCB101 PCB153 PCB138 PCB180

    (c)

    0%

    20%

    40%

    60%

    80%

    100%

    KOS KUC CHU WOL UNT GRU

    PeCB HCB α-HCH β-HCH γ-HCH δ-HCHo,p'-DDE p,p'-DDE o,p'-DDD p,p'-DDD o,p'-DDT p,p'-DDT

    Fig. 1 a PAH, b PCB and c OCPpatterns across sites

    23434 Environ Sci Pollut Res (2019) 26:23429–23441

  • average for 31%. This may reflect the different land use of thepast, given that DDTwas used for agricultural purposes in thisregion of the Czech Republic until the 1970s.

    Seasonal and spatial variations

    Significantly (p < 0.05) higher PAH deposition fluxes wereobserved at KUC than at the other sites. The spatial variationacross the other sites, apart from KUC, was lower for PAHsthan for PCBs. The maximum at the rural site KUC was ex-pected, because of local emissions, namely domestic heatingfrom a nearby village with road traffic and agricultural ma-chinery, unlike at the other sites. The PAH deposition flux atKUC site ranged from 49 to 1100 ng m−2 d−1 with an averageof 140 ngm−2 d−1, i.e. 2–6 times higher than for the other sites(Fig. 2, Table S4). The highest flux of PAHs at KUC wasobserved in spring-summer, while at KOS, it was highest inautumn–winter and at WOL it was highest in summer. Theseresults are somewhat unexpected, because the concentrationof PAHs in the air are higher in winter everywhere (Dat andChang 2017), which results from the combination of severalfactors i.e. increased sources (i.e. domestic heating), meteoro-logical conditions (i.e. lower height of mixed layer; Finlayson-Pitts and Pitts 2000; Dvorská et al. 2012) and longer photo-chemical lifetime (Finlayson-Pitts and Pitts 2000; Dvorskáet al. 2012). Obviously, the seasonal variations of PAH bulkdeposition is not dominated by atmospheric concentrationthroughout the region, but influenced by other, spatially vari-able factors, such as meteorological. Also, earlier studies(Dickhut and Gustafson 1995; Fernández et al. 2003) ex-plained the seasonality of the bulk deposition flux by the sea-sonality of precipitation. In the here studied region, highestprecipitation was generally observed in summer (Table S1).Most of the studies (Halsall et al. 1997; Ollivon et al. 2002;Blanchard et al. 2007; Gocht et al. 2007; Esen et al. 2008;Birgül et al. 2011; Binici et al. 2014) reported highest levelsof PAHs flux in winter. Moreover, we note that the influenceof open fires on the PAH time series cannot be excluded.

    The PAH bulk deposition time series in KOS, ≈4 years,does not show a downward trend (Table S4, Fig. 2). A down-ward trend could reflect and would eventually confirm ongo-ing mitigation measures (EEA 2017). Because of interannualvariation of precipitation and other environmental parameters,much longer time series would be needed to identify a signif-icant long-term trend.

    Significantly (p < 0.05) higher deposition flux of Σ6PCBwas found at KUC, where it ranged from 64 to 4400 pg m−2

    d−1 with an average of 140 pg m−2 d−1, i.e. which is 1.4–4.5times higher than for the other sites (Table S4). This wasexpected because of the rather intense historical use of PCBsin this region (electrical equipment and constructions;Christan and Janse 2005). Regarding the seasonal variations,a higher flux of Σ6PCBs was generally measured in summer

    in KOS, KUC and WOL, while the lowest mass flux to allsites was generally measured in autumn. At the other sites, noobvious seasonal variation was found. There is a peak in sum-mer 2012 in KOS. The possible reason is the construction of ameteorological tower at the observatory, distanced of about100 m from the sampling site that may have enhancedrevolatilisation from soils. PCB deposition fluxes were report-ed with varying seasonality from sites across Europe(Brorström-Lundén et al. 1994; Agrell et al. 2002; Carreraet al. 2002; Teil et al. 2004; Nizzetto et al. 2006; Blanchardet al. 2007; Bergknut et al. 2011; Newton et al. 2014).

    The total deposition fluxes for Σ12OCPs observed for thedifferent sites (Fig. 2, Table S4) were not statistically differentfrom each other. This supports the perception of long-lived(persistent), regionally distributed pollutants. However, therewas one exception, namely PeCB was much higher at theAustrian sites (averages ranging 110–250 pg m−2 d−1) thanat the Czech sites (averages ranging 12–29 pg m−2 d−1, sametime period; Fig. 1c). The reason is unknown.

    The highest flux of Σ12OCPs was generally measured insummer, but without significant seasonal variation. Jakobiet al. (2015) reported higher deposition fluxes in summer forHCB and HCH in Central Europe, but obvious differencesacross sites for DDX compounds, similar to this study. Thevariability was attributed to different origin of air massesadvected to the sites. Arellano et al. (2015) also reportedhigher flux of HCH in spring–summer in Slovakia. Newtonet al. (2014) did not find an obvious seasonal variation.

    The ratio α-HCH/γ-HCH was not significantly differentacross sites (p < 0.05) and smaller than 1 in all seasons at allsites. This suggests using lindane rather than technical HCH.The fraction of 5 ring PAHs among all measured PAHs wasnot significantly different (p < 0.05) across sites with oneexception, comparing GRU and WOL. These two ratios sug-gest that pollutants are distributed equally in Central Europe.The ratio between sum of all DDT isomers over all isomers ofDDX compounds was significantly different (p < 0.05) be-tween KOS and all other sites with one exception, GRU.This ratio was highest at KOS, suggesting most recent usageof DDT in that area. The ratio between o,p isomers and allisomers of DDX compounds was significantly different (p <0.05) between KOS and all sites, also, between KUC andCHU, GRU and then between UNT and GRU. These resultsmust be viewed with caution for the Austrian sites, because ofthe low number (N = 4) of samples.

    Distribution between XAD and GFF

    The sampler used was designed such that the freely dissolvedphase was collected on XAD while the particulate phase wascollected on GFF. Therefore, the fraction observed on GFFwill be described using the particulate fraction from deposi-tion (θdep), which should not be confused with the particulate

    Environ Sci Pollut Res (2019) 26:23429–23441 23435

  • mass fraction in aerosols (θ). It is known that for apolar ormid-polar compounds particle, scavenging is more efficientthan gas scavenging (Bidleman 1988); therefore, we can

    expect that θdep will be higher than θ. However, we cannotexclude sampling artefacts affecting the accuracy of θdep, asdiscussed above (BMaterial and methods^).

    (a)

    (b)

    (c)

    1

    10

    100

    1000

    10000

    Fb

    ulk

    (ng d

    ay-1

    m-2

    )

    KOS KUC CHU WOL UNT GRU

    1

    10

    100

    1000

    10000

    Fb

    ulk

    (ng d

    ay-1

    m-2

    )

    KOS KUC CHU WOL UNT GRU

    1

    10

    100

    1000

    10000

    Fb

    ulk

    (ng d

    ay-1

    m-2

    )

    KOS KUC CHU WOL UNT GRU

    Fig. 2 Bulk deposition fluxes of a 15 PAHs, b 6 PCBs and c 12 OCPs shown in log axis

    23436 Environ Sci Pollut Res (2019) 26:23429–23441

  • (a)

    (b)

    (c)

    3042 41 37

    49 47 4858 53 59 61 62

    70 64 69

    7058 59 63

    51 53 5242 47 41 39 38

    30 36 31

    0%

    20%

    40%

    60%

    80%

    100%

    ACY ACE FLN PHE ANT FLT PYR BAA CHR BBF BKF BAP IPY DHA BPE

    GFF XAD

    30 3856 60

    73 74

    70 6244 40

    27 26

    0%

    20%

    40%

    60%

    80%

    100%

    PCB28 PCB52 PCB101 PCB153 PCB138 PCB180

    GFF XAD

    36 26 15 15 15 18 2555 55 48 57 59

    64 74 85 85 85 82 7545 45 52 43 41

    0%

    20%

    40%

    60%

    80%

    100%

    GFF XADFig. 3 Average distribution between GFF and XAD of a PAHs, b PCBs and c OCPs

    Environ Sci Pollut Res (2019) 26:23429–23441 23437

  • Only measurements for which SOCs were > iLOQ in bothGFF and XAD are considered. For PAHs and PCBs, θdepgenerally increased with increasing molecular weight and/ordecreasing volatility (Fig. 3). However, this was not observedfor OCPs (Fig. 3). Moreover, θdep was generally higher insummer than in winter for all groups of compounds(Figure S3). This implies that the particulate mass fraction inaerosols, θ, which shows the opposite seasonality (e.g. forPAHs and OCPs in the study region: Shahpoury et al. 2015;Degrendele et al. 2014), is not preserved in total deposition.The same shift was already reported for wet deposition fluxesfrom KOS (Škrdlíková et al. 2011). The seasonal variations ofcloud depth and cloud base height, and of the scavengingefficiencies of rain and snow for gases and particles(Bidleman 1988; Škrdlíková et al. 2011; Wania andWestgate 2008), as well as of wind velocity, influencing dryparticle deposition (Pryor et al. 2007), may well explain thefinding.

    Apart from environmental parameters, sampling artefactssuch as temperature-driven volatilisation from the surface ofthe sampler, stronger in summer, could have shifted θdep asobserved. Also, less precipitation in winter (Table S1) mayhave caused a negative sampling artefact (i.e. collected partic-ulate material not efficiently washed from the surface of thesampler to the sampling media), eventually contributed toshifted θdep as observed.

    Conclusions

    We studied atmospheric bulk deposition of a number of or-ganic pollutants at 6 sites in Central Europe during 2011–2015. The time series (1–≈ 4 years) are too short, to addresslong-term trends. The results suggest that atmospheric depo-sition in 2010 is an important pathway of pollution transfer toecosystems in the Central European background. The sub-stance patterns are quite similar across sites (except for one,the rural site, which is explained by historical usage/pollutionof HCH and DDT; and except relatively higher PeCB deposi-tion at the Austrian sites).

    For substances which deposition is dominated by dry par-ticle deposition, and because of the significance of surfaceroughness for dry particle deposition (McLachlan andHorstmann 1998; Pryor et al. 2007; Glüge et al. 2015), theobserved patterns may deviate significantly from the sub-stance patterns the natural surfaces (grassland, cropland, for-est) are subject to. This also implies a systematic underesti-mate of the flux of those pollutants which mostly partition toparticulate matter in ambient aerosols. For this reason andbecause of the negative sampling artefact arising fromvolatilisation losses of dry deposited gaseous substances fromthe sampler surface, fluxes derived from bulk deposition sam-plers in use should be understood as lower estimates of the

    flux into terrestrial ecosystems. This underestimate is signifi-cant as dry deposition is more efficient than wet deposition forPAHs (Škrdlíková et al. 2011) and, considering substanceproperties (Koa, besides other; Wania and Westgate 2008),even more so for PCBs and OCPs. While the volatilisationlosses might be unavoidable for long sampling periods, thedry particle deposition efficiency could be mimicked morerealistic by a sampler design with higher surface roughness,e.g. mimicking the surface roughness of particularly forest,but also cropland or grassland. Similarly, deposition to stonefaçades had been mimicked using a sampler surface with theidentical roughness as the façade under study (Lammel andMetzig 1997).

    Acknowledgements Open access funding provided by Max PlanckSociety. We thank Ondřej Sáňka for help with geographical visualisation.

    Funding information This work was funded by the EuropeanCommission through the Territorial Cooperation Programme Austria–Czech Republic 2007–2013 (MonAirNet project, No. M00124), theMinistry of Education, Youth and Sports of the Czech Republic throughResearch Infrastructure (project LM2015051) and ACTRIS-CZ(LM2015037 and CZ.02.1.01/0.0/0.0/16_013/0001315) and by theCzech Science Foundation (GAČR 503 16-11537S).

    Compliance with ethical standards

    Conflict of interest The authors declare that they have no conflict ofinterest.

    Open Access This article is distributed under the terms of the CreativeCommons At t r ibut ion 4 .0 In te rna t ional License (h t tp : / /creativecommons.org/licenses/by/4.0/), which permits unrestricted use,distribution, and reproduction in any medium, provided you giveappropriate credit to the original author(s) and the source, provide a linkto the Creative Commons license, and indicate if changes were made.

    References

    Agrell C, Larsson P, Okla L, Agrell J (2002) PCB congeners in precipi-tation, wash out ratios and depositional fluxes within the Baltic Searegion, Europe. Atmos Environ 36:371–383. https://doi.org/10.1016/S1352-2310(01)00228-X

    Ali U, Syed JH, Malik RN, Katsoyiannis A, Li J, Zhang G, Jones KC(2014) Organochlorine pesticides (OCPs) in South Asian region: areview. Sci Total Environ 476–477:705–717. https://doi.org/10.1016/j.scitotenv.2013.12.107

    AmodioM, deGennaro G, di Gilio A, TutinoM (2014)Monitoring of thedeposition of PAHs and metals produced by a steel plant in Taranto(Italy). Adv Meteorol 2014:1–10. https://doi.org/10.1155/2014/598301

    Arellano L, Fernández P, Fonts R, Rose NL, Nickus U, Thies H, StuchlíkE, Camarero L, Catalan J, Grimalt JO (2015) Increasing and de-creasing trends of the atmospheric deposition of organochlorinecompounds in European remote areas during the last decade.Atmos Chem Phys 15:6069–6085. https://doi.org/10.5194/acp-15-6069-2015

    23438 Environ Sci Pollut Res (2019) 26:23429–23441

    https://doi.org/10.1016/S1352-2310(01)00228-Xhttps://doi.org/10.1016/S1352-2310(01)00228-Xhttps://doi.org/10.1016/j.scitotenv.2013.12.107https://doi.org/10.1016/j.scitotenv.2013.12.107https://doi.org/10.1155/2014/598301https://doi.org/10.1155/2014/598301https://doi.org/10.5194/acp-15-6069-2015https://doi.org/10.5194/acp-15-6069-2015

  • Atlas E, Giam CS (1988) Ambient concentration and precipitation scav-enging of atmospheric organic pollutants. Water Air Soil Pollut 38:19–36. https://doi.org/10.1007/BF00279583

    Backe C, Larsson P, Agrell C (2002) Spatial and temporal variation ofpolychlorinated biphenyl (PCB) in precipitation in southernSweden. Sci Total Environ 285:117–132. https://doi.org/10.1016/S0048-9697(01)00901-9

    Bansal V, Kim KH (2015) Review of PAH contamination in food prod-ucts and their health hazards. Environ Int 84:26–38. https://doi.org/10.1016/j.envint.2015.06.016

    Bergknut M, Laudon H, Jansson S, Larsson A, Gocht T, Wiberg K (2011)Atmospheric deposition, retention, and stream export of dioxins andPCBs in a pristine boreal catchment. Environ Pollut 159:1592–1598. https://doi.org/10.1016/j.envpol.2011.02.050

    Bidleman TF (1988) Atmospheric processes. Environ Sci Technol 22:361–367. https://doi.org/10.1021/es00169a002

    Binici B, Yenisoy-Karakaş S, Bilsel M, Durmaz-Hilmioglu N (2014)Sources of polycyclic hydrocarbons and pesticides in soluble frac-tion of deposition samples in Kocaeli, Turkey. Environ Sci PollutRes 21:2907–2917. https://doi.org/10.1007/s11356-013-2239-z

    Birgül A, Tasdemir Y, Cindoruk SS (2011) Atmospheric wet and drydeposition of polycyclic aromatic hydrocarbons (PAHs) determinedusing a modified sampler. Atmos Res 101:341–353. https://doi.org/10.1016/j.atmosres.2011.03.012

    Blanchard M, Teil MJ, Guigon E, Larcher-Tiphagne K, Ollivon D,Garban B, Chevreuil M (2007) Persistent toxic substance inputs tothe river Seine basin (France) via atmospheric deposition and urbansludge application. Sci Total Environ 375:232–243. https://doi.org/10.1016/j.scitotenv.2006.12.012

    Brorström-Lundén E, Lindskog A, Mowrer J (1994) Concentrations andfluxes of organic compounds in the atmosphere of the Swedish westcoast. Atmos Environ 28:3605–3615. https://doi.org/10.1016/1352-2310(94)00194-P

    Carrera G, Fernández P, Grimalt JO, Ventura M, Camarero L, Catalan J,Nickus U, Thies H, Psenner R (2002) Atmospheric deposition oforganochlorine compounds to remote high mountain lakes ofEurope. Environ Sci Technol 36:2581–2588. https://doi.org/10.1021/es0102585

    Castro-Jiménez J, Dachs J, Eisenreich SJ (2015) Atmospheric depositionof POPs: Implications for the chemical pollution of aquatic environ-ments. In: Zeng EY, Barcelo D (eds) Persistent Organic Pollutants(POPs): Analytical Techniques, Environmental Fate and BiologicalEffects, Chapter 8, vol 67. Elsevier - Comprehensive AnalyticalChemistry, pp 295–322. https://doi.org/10.1016/B978-0-444-63299-9.00008-9

    Cetin B, Öztürk F, Keles M, Yurdakul S (2017) PAHs and PCBs in anEastern Mediterranean megacity, Istanbul: their spatial and temporaldistributions, air-soil exchange and toxicological effects. EnvironPollut 220:1322–1332. https://doi.org/10.1016/j.envpol.2016.11.002

    Christan E, Janse J (2005) EuroPCB: inventory PCB enforcement inmember states. Part II: Fiches - Results for each member state.Inspectorate of the Ministry of Housing, Spatial Planning and theEnvironment South Unit of the Netherlands, vol 68, Vienna http://www.cleen-europe.eu/file/download/71/EuroPCB_part_II_fiches_final.pdf. Accessed 4 Feb 2019

    Cindoruk SS, Tasdemir Y (2014) The investigation of atmospheric depo-sition distribution of organochlorine pesticides (OCPs) in Turkey.Atmos Environ 87:207–217. https://doi.org/10.1016/j.atmosenv.2014.01.008

    Čupr P, Pěnkava B (2012) Vzorkovač atmosférické depozice(Atmospheric deposition sampler). Patent. No. 23347. (owner:Masarykova univerzita, CZ, BAGHIRRA s.r.o. Praha, CZ).Industrial Property Office, Czech Republic

    Dat ND, Chang MB (2017) Review on characteristics of PAHs in atmo-sphere, anthropogenic sources and control technologies. Sci Total

    Environ 609:682–693. https://doi.org/10.1016/j.scitotenv.2017.07.204

    Degrendele C, Okonski K, Melymuk L, Landlová L, Kukučka P, Čupr P,Klánová J (2014) Size specific distribution of the atmospheric par-ticulate PCDD/Fs, dl-PCBs and PAHs on a seasonal scale: implica-tions for cancer risks from inhalation. Atmos Environ 98:410–416.https://doi.org/10.1016/j.atmosenv.2014.09.001

    Dickhut RM, Gustafson KE (1995) Atmospheric washout of polycyclicaromatic hydrocarbons in the Southern Chesapeake Bay region.Environ Sci Technol 29:1518–1525. https://doi.org/10.1021/es00006a013

    DIN (2002) German Industrial Standard. Luftbeschaffenheit undBodenbeschaffenheit - Messen der atmosphärischen Depositionorganischer Spurenstoffe; trichter-adsorber-verfahren - Teil 1:Sammelgeräte; Anforderungen, Aufbau, Anwendung. In: DIN, vol19739-1. Deutsches Institut für Normung e.V., Berlin, pp 1–17

    Dvorská A, Komprdová K, Lammel G, Klánová J, Plachá H (2012)Polycyclic aromatic hydrocarbons in background air in centralEurope - seasonal levels and limitations for source apportionment.Atmos Environ 46:147–154. https://doi.org/10.1016/j.atmosenv.2011.10.007

    EEA (2017) European Union Emission Inventory Report 1990–2016under the UNECE Convention on Long-Range Transboundary AirPollution (LRTAP).

    el Shahawi MS, Hamza A, Bashammakh AS, al Saggaf WT (2010) Anoverview on the accumulation, distribution, transformations, toxic-ity and analytical methods for the monitoring of persistent organicpollutants. Talanta 80:1587–1597. https://doi.org/10.1016/j.talanta.2009.09.055

    Esen F, Cindoruk SS, Tasdemir Y (2008) Bulk deposition of polycyclicaromatic hydrocarbons (PAHs) in an industrial site of Turkey.Environ Pollut 152:461–467. https://doi.org/10.1016/j.envpol.2007.05.031

    Feng D, Liu Y, Gao Y, Zhou J, Zheng L, Qiao G, Ma L, Lin Z, GrathwohlP (2017) Atmospheric bulk deposition of polycyclic aromatic hy-drocarbons in Shanghai: temporal and spatial variation, and globalcomparison. Environ Pollut 230:639–647. https://doi.org/10.1016/j.envpol.2017.07.022

    Fernández P, Carrera G, Grimalt JO, Ventura M, Camarero L, Catalan J,Nickus U, Thies H, Psenner R (2003) Factors governing the atmo-spheric deposition of polycyclic aromatic hydrocarbons to remoteareas. Environ Sci Technol 37:3261–3267. https://doi.org/10.1021/es020137k

    Finlayson-Pitts BJ, Pitts JN (2000) Chemistry of the upper and loweratmosphere. Academic Press, San Diego. https://doi.org/10.1016/B978-012257060-5/50010-1

    Foan L, Domercq M, Bermejo R, Santamaria JM, Simon V (2012)Polycyclic aromatic hydrocarbons (PAHs) in remote bulk andthroughfall deposition: seasonal and spatial trends. Environ EngManag J 11:1101–1110. https://doi.org/10.30638/eemj.2012.134

    Franz TP, Eisenreich SJ, Swanson MB (1991) Evaluation of precipitationsamplers for assessing atmospheric fluxes of trace organic contam-inants. Chemosphere 23:343–361. https://doi.org/10.1016/0045-6535(91)90189-K

    Gambaro A, Radaelli M, Piazza R, Stortini AM, Contini D, Belosi F,Zangrando R, Cescon P (2009) Organic micropollutants in wetand dry depositions in the Venice Lagoon. Chemosphere 76:1017–1022. https://doi.org/10.1016/j.chemosphere.2009.04.063

    Glüge J, Bogdal C, Scheringer M, Hungerbühler K (2015) Atmosphericgas-particle partitioning versus gaseous/particle-bound deposition ofSVOCs: why they are not equivalent. Atmos Environ 115:317–324.https://doi.org/10.1016/j.atmosenv.2015.05.028

    Gocht T, Klemm O, Grathwohl P (2007) Long-term atmospheric bulkdeposition of polycyclic aromatic hydrocarbons (PAHs) in ruralareas of Southern Germany. Atmos Environ 41:1315–1327.https://doi.org/10.1016/j.atmosenv.2006.09.036

    Environ Sci Pollut Res (2019) 26:23429–23441 23439

    https://doi.org/10.1007/BF00279583https://doi.org/10.1016/S0048-9697(01)00901-9https://doi.org/10.1016/S0048-9697(01)00901-9https://doi.org/10.1016/j.envint.2015.06.016https://doi.org/10.1016/j.envint.2015.06.016https://doi.org/10.1016/j.envpol.2011.02.050https://doi.org/10.1021/es00169a002https://doi.org/10.1007/s11356-013-2239-zhttps://doi.org/10.1016/j.atmosres.2011.03.012https://doi.org/10.1016/j.atmosres.2011.03.012https://doi.org/10.1016/j.scitotenv.2006.12.012https://doi.org/10.1016/j.scitotenv.2006.12.012https://doi.org/10.1016/1352-2310(94)00194-Phttps://doi.org/10.1016/1352-2310(94)00194-Phttps://doi.org/10.1021/es0102585https://doi.org/10.1021/es0102585https://doi.org/10.1016/B978-0-444-63299-9.00008-9https://doi.org/10.1016/B978-0-444-63299-9.00008-9https://doi.org/10.1016/j.envpol.2016.11.002https://doi.org/10.1016/j.envpol.2016.11.002http://www.cleen-europe.eu/file/download/71/EuroPCB_part_II_fiches_%20final.pdfhttp://www.cleen-europe.eu/file/download/71/EuroPCB_part_II_fiches_%20final.pdfhttp://www.cleen-europe.eu/file/download/71/EuroPCB_part_II_fiches_%20final.pdfhttps://doi.org/10.1016/j.atmosenv.2014.01.008https://doi.org/10.1016/j.atmosenv.2014.01.008https://doi.org/10.1016/j.scitotenv.2017.07.204https://doi.org/10.1016/j.scitotenv.2017.07.204https://doi.org/10.1016/j.atmosenv.2014.09.001https://doi.org/10.1021/es00006a013https://doi.org/10.1021/es00006a013https://doi.org/10.1016/j.atmosenv.2011.10.007https://doi.org/10.1016/j.atmosenv.2011.10.007https://doi.org/10.1016/j.talanta.2009.09.055https://doi.org/10.1016/j.talanta.2009.09.055https://doi.org/10.1016/j.envpol.2007.05.031https://doi.org/10.1016/j.envpol.2007.05.031https://doi.org/10.1016/j.envpol.2017.07.022https://doi.org/10.1016/j.envpol.2017.07.022https://doi.org/10.1021/es020137khttps://doi.org/10.1021/es020137khttps://doi.org/10.1016/B978-012257060-5/50010-1https://doi.org/10.1016/B978-012257060-5/50010-1https://doi.org/10.30638/eemj.2012.134https://doi.org/10.1016/0045-6535(91)90189-Khttps://doi.org/10.1016/0045-6535(91)90189-Khttps://doi.org/10.1016/j.chemosphere.2009.04.063https://doi.org/10.1016/j.atmosenv.2015.05.028https://doi.org/10.1016/j.atmosenv.2006.09.036

  • González-Gaya B, Fernández-Pinos MC, Morales L, Méjanelle L, AbadE, Piña B, Duarte CM, Jiménez B, Dachs J (2016) High atmosphere-ocean exchange of semivolatile aromatic hydrocarbons. Nat Geosci9:438–442. https://doi.org/10.1038/ngeo2714

    Halsall CJ, Coleman PJ, Jones KC (1997) Atmospheric deposition ofpolychlorinated dibenzo-p-dioxins/dibenzofurans (PCDD/Fs) andpolycyclic aromatic hydrocarbons (PAHs) in two UK cities.Chemosphere 35:1919–1931. https://doi.org/10.1016/S0045-6535(97)00265-8

    Hansson K, Palm Cousins A, Brorström-Lundén E (2006) Atmosphericconcentrations in air and deposition fluxes of POPs at Råö andPallas, trends and seasonal and spatial variations. http://urn.kb.se/resolve?urn = urn:nbn:se:naturvardsverket:diva-914 (accessedFeb 5, 2019)

    Holoubek I, Klánová J, Jarkovský J, Kohoutek J (2007) Trends in back-ground levels of persistent organic pollutants at Kosetice observato-ry, Czech Republic. Part I. Ambient air and wet deposition 1996-2005. J Environ Monit 9:557–563. https://doi.org/10.1039/b700750g

    Horstmann M, McLachlan MS (1997) Sampling bulk deposition ofpolychlorinated dibenzo-p-dioxins and dibenzofurans. AtmosEnviron 31:2977–2982. https://doi.org/10.1016/S1352-2310(97)00106-4

    IARC (2010) Some non-heterocyclic polycyclic aromatic hydrocarbonsand some related exposures. IARC Monographs on the Evaluationof Carcinogenic Risks to Humans Vol. 92. World HealthOrganisation International Agency for Research on Cancer, Lyon,p 868. https://doi.org/10.1002/14356007.a04

    Jakobi G, Kirchner M, Henkelmann B, Körner W, Offenthaler I, MocheW, Weiss P, Schaub M, Schramm KW (2015) Atmospheric bulkdeposition measurements of organochlorine pesticides at three al-pine summits. Atmos Environ 101:158–165. https://doi.org/10.1016/j.atmosenv.2014.10.060

    Jantunen LM, Bidleman TF (2006) Henry’s law constants for hexachlo-robenzene, p,p′-DDE and components of technical chlordane andestimates of gas exchange for Lake Ontario. Chemosphere 62:1689–1696. https://doi.org/10.1016/j.chemosphere.2005.06.035

    Karacik B, Okay OS, Henkelmann B, Pfister G, Schramm KW (2013)Water concentrations of PAH, PCB and OCP by using semiperme-able membrane devices and sediments.Mar Pollut Bull 70:258–265.https://doi.org/10.1016/j.marpolbul.2013.02.031

    Keyte IJ, Harrison RM, Lammel G (2013) Chemical reactivity and long-range transport potential of polycyclic aromatic hydrocarbons - areview. Chem Soc Rev 42:9333–9391. https://doi.org/10.1039/C3CS60147A

    Lammel G, Metzig G (1997) Pollutant fluxes onto the façades of a his-torical monument. Atmos Environ 31:2249–2259. https://doi.org/10.1016/S1352-2310(97)00034-4

    Lammel G, Stemmler I (2012) Fractionation and current time trends ofPCB congeners: evolvement of distributions 1950-2010 studiedusing a global atmosphere-ocean general circulation model. AtmosChem Phys 12:7199–7213. https://doi.org/10.5194/acp-12-7199-2012

    Lei YD, Wania F (2004) Is rain or snow a more efficient scavenger oforganic chemicals? Atmos Environ 38:3557–3571. https://doi.org/10.1016/j.atmosenv.2004.03.039

    Ludewig G, Robertson LW (2013) Polychlorinated biphenyls (PCBs) asinitiating agents in hepatocellular carcinoma. Cancer Lett 334:46–55. https://doi.org/10.1016/j.canlet.2012.11.041

    McLachlan MS (1998) Basic consideration and practical experience inthe development of methods to measure atmospheric deposition ofpersistent organic pollutants. In: Lükewille A (ed) EuropeanMonitoring and Evaluation Programme, Report # EMEP/CCC8/98EMEP experts meeting on measurements of persistent organicpollutants (POPs) in air and precipitation. Lillehammer, Norway, pp87–101

    McLachlan MS, Horstmann M (1998) Forests as filters of airborne or-ganic pollutants: a model. Environ Sci Technol 32:413–420. https://doi.org/10.1021/es970592u

    Meijer SN, Dachs J, Fernandez P, Camarero L, Catalan J, Del Vento S,van Drooge B, Jurado E, Grimalt JO (2006) Modelling the dynamicair–water–sediment coupled fluxes and occurrence ofpolychlorinated biphenyls in a high altitude lake. Environ Pollut140:546–560. https://doi.org/10.1016/j.envpol.2005.06.015

    Newton S, Bidleman T, Bergknut M, Racine J, Laudon H, Giesler R,Wiberg K (2014) Atmospheric deposition of persistent organic pol-lutants and chemicals of emerging concern at two sites in northernSweden. Environ Sci Process Impacts 16:298–305. https://doi.org/10.1039/c3em00590a

    Nizzetto L, Cassani C, di Guardo A (2006) Deposition of PCBs in moun-tains: the forest filter effect of different forest ecosystem types.Ecotoxicol Environ Saf 63:75–83. https://doi.org/10.1016/j.ecoenv.2005.05.005

    Offenthaler I, Jakobi G, Kaiser A, Kirchner M, Krauchi N, NiedermoserB, Schramm KW, Sedivy I, Staudinger M, Thanner G, Weiss P,Moche W (2009) Novel sampling methods for atmospheric semi-volatile organic compounds (SOCs) in a high altitude alpine envi-ronment. Environ Pollut 157:3290–3297. https://doi.org/10.1016/j.envpol.2009.05.053

    Ollivon D, Blanchoud H, Motélay-Massei A, Garban B (2002)Atmospheric deposition of PAHs to an urban site, Paris, France.Atmos Environ 36:2891–2900. https://doi.org/10.1016/S1352-2310(02)00089-4

    Pryor SC, Larsen SE, Sorensen LL, Barthelmie RJ, Grönholm T, KulmalaM, Launiainen S, Rannik U, Vesala T (2007) Particle fluxes overforests: analyses of flux methods and functional dependencies. JGeophys Res 112. https://doi.org/10.1029/2006JD008066

    Qu C, Albanese S, Lima A, Hope D, Pond P, Fortelli A, Romano N,Cerino P, Pizzolante A, De Vivo B (2019) The occurrence ofOCPs, PCBs, and PAHs in the soil, air, and bulk deposition of theNaples metropolitan area, southern Italy: implications for sourcesand environmental processes. Environ Int 124:89–97. https://doi.org/10.1016/j.envint.2018.12.031

    Ross G (2004) The public health implications of polychlorinated biphe-nyls (PCBs) in the environment. Ecotoxicol Environ Saf 59:275–291. https://doi.org/10.1016/j.ecoenv.2004.06.003

    Rossini P, Matteucci G, Raccanelli S, Favotto M, Guerzoni S, Gattolin M(2007) Polycyclic aromatic hydrocarbons in atmospheric deposi-tions around the Venice Lagoon. Polycycl Aromat Compd 27:197–210. https://doi.org/10.1080/10406630701359773

    Scheringer M, Stroebe M, Wania F, Wegmann F, Hungerbühler K (2004)The effect of export to the deep sea on the long-range transportpotential of persistent organic pollutants. Environ Sci Pollut Res11:41–48. https://doi.org/10.1065/espr2003.11.176

    Schifman LA, Boving TB (2015) Spatial and seasonal atmospheric PAHdeposition patterns and sources in Rhode Island. Atmos Environ120:253–261. https://doi.org/10.1016/j.atmosenv.2015.08.056

    Shahpoury P, Lammel G, Holubová Šmejkalová A, Klánová J, PřibylováP, Váňa M (2015) Polycyclic aromatic hydrocarbons,polychlorinated biphenyls, and chlorinated pesticides in backgroundair in central Europe - investigating parameters affecting wet scav-enging of polycyclic aromatic hydrocarbons. Atmos Chem Phys 15:1795–1805. https://doi.org/10.5194/acp-15-1795-2015

    Shen L, Wania F (2005) Compilation, evaluation, and selection ofphysical-chemical property data for organochlorine pesticides. JChem Eng Data 50:742–768. https://doi.org/10.1021/je049693f

    Škrdlíková L, Landlová L, Klánová J, Lammel G (2011) Wet depositionand scavenging efficiency of gaseous and particulate phase polycy-clic aromatic compounds at a central European suburban site. AtmosEnviron 45:4305–4312. https://doi.org/10.1016/j.atmosenv.2011.04.072

    23440 Environ Sci Pollut Res (2019) 26:23429–23441

    https://doi.org/10.1038/ngeo2714https://doi.org/10.1016/S0045-6535(97)00265-8https://doi.org/10.1016/S0045-6535(97)00265-8http://urn.kb.se/resolve?urnhttp://urn.kb.se/resolve?urnhttps://doi.org/10.1039/b700750ghttps://doi.org/10.1039/b700750ghttps://doi.org/10.1016/S1352-2310(97)00106-4https://doi.org/10.1016/S1352-2310(97)00106-4https://doi.org/10.1002/14356007.a04https://doi.org/10.1016/j.atmosenv.2014.10.060https://doi.org/10.1016/j.atmosenv.2014.10.060https://doi.org/10.1016/j.chemosphere.2005.06.035https://doi.org/10.1016/j.marpolbul.2013.02.031https://doi.org/10.1039/C3CS60147Ahttps://doi.org/10.1039/C3CS60147Ahttps://doi.org/10.1016/S1352-2310(97)00034-4https://doi.org/10.1016/S1352-2310(97)00034-4https://doi.org/10.5194/acp-12-7199-2012https://doi.org/10.5194/acp-12-7199-2012https://doi.org/10.1016/j.atmosenv.2004.03.039https://doi.org/10.1016/j.atmosenv.2004.03.039https://doi.org/10.1016/j.canlet.2012.11.041https://doi.org/10.1021/es970592uhttps://doi.org/10.1021/es970592uhttps://doi.org/10.1016/j.envpol.2005.06.015https://doi.org/10.1039/c3em00590ahttps://doi.org/10.1039/c3em00590ahttps://doi.org/10.1016/j.ecoenv.2005.05.005https://doi.org/10.1016/j.ecoenv.2005.05.005https://doi.org/10.1016/j.envpol.2009.05.053https://doi.org/10.1016/j.envpol.2009.05.053https://doi.org/10.1016/S1352-2310(02)00089-4https://doi.org/10.1016/S1352-2310(02)00089-4https://doi.org/10.1029/2006JD008066https://doi.org/10.1016/j.envint.2018.12.031https://doi.org/10.1016/j.envint.2018.12.031https://doi.org/10.1016/j.ecoenv.2004.06.003https://doi.org/10.1080/10406630701359773https://doi.org/10.1065/espr2003.11.176https://doi.org/10.1016/j.atmosenv.2015.08.056https://doi.org/10.5194/acp-15-1795-2015https://doi.org/10.1021/je049693fhttps://doi.org/10.1016/j.atmosenv.2011.04.072https://doi.org/10.1016/j.atmosenv.2011.04.072

  • Staelens J, de Schrijver A, van Avermaet P, Genouw G, Verhoest N(2005) A comparison of bulk and wet-only deposition at two adja-cent sites in Melle (Belgium). Atmos Environ 39:7–15. https://doi.org/10.1016/j.atmosenv.2004.09.055

    Stemmler I, Lammel G (2012) Long-term trends of continental-scalePCB patterns studied using a global atmosphere-ocean general cir-culation model. Environ Sci Pollut Res 19:1971–1980. https://doi.org/10.1007/s11356-012-0943-8

    Teil MJ, Blanchard M, Chevreuil M (2004) Atmospheric deposition oforganochlorines (PCBs and pesticides) in northern France.Chemosphere 55:501–514. https://doi.org/10.1016/j.chemosphere.2003.11.064

    UNECE (1999) Economic Commission for Europe. 1979 Convention onlong-range transboundary air pollution and its 1998 protocols onpersistent organic pollutants and heavy metals. United Nations,New York, 7 pp, 49 pp and 33 pp

    UNEP (2003) United Nations Environment Programme. Regionallybased assessment of persistent toxic substances, Global report2003. United Nations, Geneva, 207 pp

    UNEP (2008) United Nations Environment Programme. StockholmConvention. http://chm.pops.int/Portals/0/download.aspx?d =UNEP-POPS-COP-CONVTEXT-2017.English.pdf (accessedApril 4, 2019)

    Wania F, Westgate JN (2008) On the mechanism of mountain cold-trapping of organic chemicals. Environ Sci Technol 42:9092–9098. https://doi.org/10.1021/es8013198

    Wania F, Hoff JT, Jia CQ, Mackay D (1998) The effects of snow and iceon the environmental behaviour of hydrophobic organic chemicals.Environ Pollut 102:25–41. https://doi.org/10.1016/S0269-7491(98)00073-6

    Wania F, Shen L, Lei YD, Teixeira C, Muir DCG (2003) Developmentand calibration of a resin-based passive sampling system for moni-toring persistent organic pollutants in the atmosphere. Environ SciTechnol 37:1352–1359. https://doi.org/10.1021/es026166c

    Xiao H, Li NQ,Wania F (2004) Compilation, evaluation, and selection ofphysical-chemical property data for alpha-, beta-, and gamma-hexa-chlorocyclohexane. J Chem Eng Data 49:173–185. https://doi.org/10.1021/je034214i

    Publisher’s note Springer Nature remains neutral with regard tojurisdictional claims in published maps and institutional affiliations.

    Environ Sci Pollut Res (2019) 26:23429–23441 23441

    https://doi.org/10.1016/j.atmosenv.2004.09.055https://doi.org/10.1016/j.atmosenv.2004.09.055https://doi.org/10.1007/s11356-012-0943-8https://doi.org/10.1007/s11356-012-0943-8https://doi.org/10.1016/j.chemosphere.2003.11.064https://doi.org/10.1016/j.chemosphere.2003.11.064http://chm.pops.int/Portals/0/download.aspx?dhttps://doi.org/10.1021/es8013198https://doi.org/10.1016/S0269-7491(98)00073-6https://doi.org/10.1016/S0269-7491(98)00073-6https://doi.org/10.1021/es026166chttps://doi.org/10.1021/je034214ihttps://doi.org/10.1021/je034214i

    Bulk atmospheric deposition of persistent organic pollutants and polycyclic aromatic hydrocarbons in Central EuropeAbstractIntroductionMaterial and methodsSamplingSample preparation and analysisQA-QC

    Results and discussionAtmospheric bulk deposition flux of PAHs, PCBs and OCPs and their composition profilesSeasonal and spatial variationsDistribution between XAD and GFF

    ConclusionsReferences